Heterogeneous or Homogeneous Catalysis? - American Chemical


Heterogeneous or Homogeneous Catalysis? - American Chemical...

25 downloads 82 Views 179KB Size

Published on Web 07/17/2004

Heterogeneous or Homogeneous Catalysis? Mechanistic Studies of the Rhodium-Catalyzed Dehydrocoupling of Amine-Borane and Phosphine-Borane Adducts Cory A. Jaska and Ian Manners* Contribution from the Department of Chemistry, UniVersity of Toronto, 80 St. George Street, Toronto, Ontario, Canada M5S 3H6 Received April 14, 2004; E-mail: [email protected]

Abstract: In depth, comparative studies on the catalytic dehydrocoupling of the amine-borane adduct Me2NH‚BH3 (to form [Me2N-BH2]2) and the phosphine-borane adduct Ph2PH‚BH3 (to form Ph2PH-BH2PPh2-BH3) with a variety of Rh (pre)catalysts such as [{Rh(1,5-cod)(µ-Cl)}2], Rh/Al2O3, Rhcolloid/[Oct4N]Cl, and [Rh(1,5-cod)2]OTf have been performed in order to determine whether the dehydrocoupling proceeds by a homogeneous or heterogeneous mechanism. The results obtained suggest that the catalytic dehydrocoupling of Me2NH‚BH3 is heterogeneous in nature involving Rh(0) colloids, while that of Ph2PH‚BH3 proceeds by a homogeneous mechanism even when starting with Rh(0) precursors such as Rh/Al2O3. The catalytic dehydrocoupling reactions are thought to proceed by different mechanisms due to a combination of factors such as (i) the greater reducing strength of amine-borane adducts, (ii) the increased ease of dissociation of phosphine-borane adducts, and (iii) phosphine ligation and/or poisoning of active catalytic sites on metal colloids.

Introduction

Transition metal catalysis plays a profound role in organic synthesis. Drug, natural product, and polymer syntheses all utilize metal-catalyzed reactions to provide alternative or improved routes to the desired end products. Thus, the formation of new C-C, C-N, C-O, and C-H bonds can be facilitated by a ready arsenal of metal-catalyzed routes, allowing for precise control over factors such as chemo-, regio-, and stereoselectivity while being performed under mild reaction conditions with increased efficiency, minimal waste generation, and decreased energy consumption. In contrast, the use of transition metalcatalyzed processes for the preparation of inorganic molecules and polymers is relatively unexplored. The development of facile catalytic processes for the synthesis of inorganic compounds would be highly desirable, as currently available synthetic strategies are often haphazard, of limited scope, or are restricted to salt metathesis reactions. In particular, detailed mechanistic studies of the catalytic cycles would allow for the development of generalized reactions, which should help to advance this promising field.1 Recently, we have shown that the heterodehydrocoupling of phosphine-borane adducts RR′PH‚BH3 can be catalyzed by a variety of Rh complexes, providing facile routes to phosphinoborane rings, chains, and high molecular weight polymers [RR′P-BH2]n.2,3 We have recently extended this work to the catalytic dehydrocoupling of amine-borane adducts RR′NH‚BH3, which provides a mild and convenient route to cycloamino(1) (a) Tilley, T. D. Acc. Chem. Res. 1993, 26, 22. (b) Gauvin, F.; Harrod, J. F.; Woo, H. G. AdV. Organomet. Chem. 1998, 42, 363. (c) Reichl, J. A.; Berry, D. H. AdV. Organomet. Chem. 1998, 43, 197. (d) Jaska, C. A.; Bartole-Scott, A.; Manners, I. Dalton Trans. 2003, 4015. 9776

9

J. AM. CHEM. SOC. 2004, 126, 9776-9785

boranes [RR′N-BH2]2 and borazines [RN-BH]3.4 Based on the catalytic chemistry displayed by amine-borane adducts, a tandem catalytic dehydrocoupling-hydrogenation reaction involving a variety of Rh (pre)catalysts and Me2NH‚BH3 as a stoichiometric hydrogen source for the hydrogenation of alkenes at 25 °C has been recently developed.5 Interestingly, the catalytic dehydrocoupling of phosphine-borane and amine-borane adducts using the common precatalyst [{Rh(1,5-cod)(µ-Cl)}2] (cod ) cyclooctadiene) were observed to afford dramatically different reaction mixtures: a clear, dark red solution for the former but a black, opaque solution with a precipitate for the latter. These initial visual observations led us to investigate the possibility that different mechanisms (homogeneous or heterogeneous) may be operating for the two different systems.6 Distinguishing between true homogeneous catalysis and soluble or insoluble metal-particle heterogeneous catalysis is a problem that has attracted much recent attention.7 For example, many catalytic processes such as hydrosilation,8 alkene or arene (2) (a) Dorn, H.; Singh, R. A.; Massey, J. A.; Lough, A. J.; Manners, I. Angew. Chem., Int. Ed. 1999, 38, 3321. (b) Dorn, H.; Singh, R. A.; Massey, J. A.; Nelson, J. M.; Jaska, C. A.; Lough, A. J.; Manners, I. J. Am. Chem. Soc. 2000, 122, 6669. (c) Dorn, H.; Vejzovic, E.; Lough, A. J.; Manners, I. Inorg. Chem. 2001, 40, 4327. (d) Dorn, H.; Rodezno, J. M.; Brunnho¨fer, B.; Rivard, E.; Massey, J. A.; Manners, I. Macromolecules 2003, 36, 291. (3) (a) Dorn, H.; Jaska, C. A.; Singh, R. A.; Lough, A. J.; Manners, I. Chem. Commun. 2000, 1041. (b) Jaska, C. A.; Dorn, H.; Lough, A. J.; Manners, I. Chem.sEur. J. 2003, 9, 271. (4) (a) Jaska, C. A.; Temple, K.; Lough, A. J.; Manners, I. Chem. Commun. 2001, 962. (b) Jaska, C. A.; Temple, K.; Lough, A. J.; Manners, I. J. Am. Chem. Soc. 2003, 125, 9424. (5) Jaska, C. A.; Manners, I. J. Am. Chem. Soc. 2004, 126, 2698. (6) For reports of our preliminary results, see ref 4b and: Jaska, C. A.; Manners, I. J. Am. Chem. Soc. 2004, 126, 1334. (7) (a) Laine, R. M. J. Mol. Catal. 1982, 14, 137. (b) Widegren, J. A.; Finke, R. G. J. Mol. Catal. A: Chem. 2003, 198, 317. (c) Dyson, P. J. Dalton Trans. 2003, 2964. 10.1021/ja0478431 CCC: $27.50 © 2004 American Chemical Society

Rh-Catalyzed Dehydrocoupling of Borane Adducts

hydrogenation,9 ring-opening polymerization,10 and crosscoupling reactions11 have been studied in detail with respect to the presence of a homogeneous or heterogeneous mechanism. Since catalytic issues such as selectivity, activity, stability, and recovery are influenced differently for homogeneous and heterogeneous catalysts, the problem of distinguishing between the two is of key significance. In an important contribution, Finke and co-workers have developed a general approach to address this problem by performing a series of tests involving reaction kinetics, transmission electron microscopy (TEM) imaging, UV-visible spectroscopy, poisoning experiments, filtration experiments, catalyst isolation, etc.7b In this paper, we report full details of our studies in which we have utilized this approach to address the question of the potential presence of homogeneous or heterogeneous catalysis for the dehydrocoupling of the adducts Me2NH‚BH3 (eq 1) and Ph2PH‚BH3 (eq 2) using a series of different Rh (pre)catalysts such as [{Rh(1,5-cod)(µ-Cl)}2], Rh/Al2O3, [Oct4N]Cl stabilized Rh colloids (hereafter referred to as Rhcolloid/[Oct4N]Cl; Oct ) n-octyl), and [Rh(1,5-cod)2]OTf.6

Results and Discussion

“Homogeneous vs Heterogeneous” Tests Performed: For the comparative studies between Me2NH‚BH3 and Ph2PH‚BH3, the following series of tests were performed: (i) Analysis by TEM. This is commonly used as a preliminary test for the presence of metal particles in catalytically active solutions.7b However, this test suffers from the drawback that if metal particles are observed, they might not be the actual species responsible for the catalysis.8c (ii) Analysis by UV-visible spectroscopy. The presence of metal colloids in solution can be deduced using UV-visible spectroscopy. For example, 10 nm Rh colloids exhibit a continuous absorption in the visible range due to a surface plasmon resonance, with a steep rise in absorbance at short wavelengths.12 (iii) Analysis of the reaction kinetics. The presence of a sigmoidal-shaped kinetic curve strongly suggests a heterogeneous catalyst. It has been shown that slow, continuous nucleation followed by rapid autocatalytic surface growth can result in soluble monodisperse nanoclusters13 or insoluble bulk metal formation.9b It is suggested that the presence of such a (8) (a) Lewis, L. N.; Lewis, N. J. Am. Chem. Soc. 1986, 108, 7228. (b) Lewis, L. N. J. Am. Chem. Soc. 1990, 112, 5998. (c) Stein, J.; Lewis, L. N.; Gao, Y.; Scott, R. A. J. Am. Chem. Soc. 1999, 121, 3693. (d) Roy, A. K.; Taylor, R. B. J. Am. Chem. Soc. 2002, 124, 9510. (9) (a) Weddle, K. S.; Aiken, J. D., III; Finke, R. G. J. Am. Chem. Soc. 1998, 120, 5653. (b) Widegren, J. A.; Bennett, M. A.; Finke, R. G. J. Am. Chem. Soc. 2003, 125, 10301. (10) (a) Wu, X.; Neckers, D. C. Macromolecules 1999, 32, 6003. (b) Temple, K.; Ja¨kle, F.; Sheridan, J. B.; Manners, I. J. Am. Chem. Soc. 2001, 123, 1355. (11) (a) Davies, I. W.; Matty, L.; Hughes, D. L.; Reider, P. J. J. Am. Chem. Soc. 2001, 123, 10139. (b) Na, Y.; Park, S.; Han, S. B.; Han, H.; Ko, S.; Chang, S. J. Am Chem. Soc. 2004, 126, 250. (12) Creighton, J. A.; Eadon, D. G. J. Chem. Soc., Faraday Trans. 1991, 87, 3881.

ARTICLES

kinetic curve is a “compelling single piece of evidence” for the formation of a heterogeneous catalyst.7b,14 (iv) Mercury poisoning experiments. Mercury is a well-known poison of heterogeneous catalysts through the formation of an amalgam or adsorption onto the catalyst surface.15 If the addition of Hg is found to suppress catalytic activity, a heterogeneous catalyst can be assumed, given that proper control experiments were performed. (v) Fractional poisoning experiments. The addition of strongly coordinating ligands (such as PPh3) can help distinguish between homogeneous and heterogeneous catalysis.7b A heterogeneous catalyst can be completely poisoned by 1 equiv of ligand to completely poison the active site. (vi) Filtration experiments. Filtration of the reaction mixture using small pore membrane filters can distinguish between soluble and insoluble catalysts.7b If the activity is lowered upon filtration, an insoluble catalyst is assumed. Analysis of Catalytically Active Solutions by TEM: For the catalytic dehydrocoupling of Me2NH‚BH3 and Ph2PH‚BH3 using the precatalyst [{Rh(1,5-cod)(µ-Cl)}2], initial TEM images were obtained in order to test for the presence of metal particles. For Me2NH‚BH3, analysis of the catalytically active solution by TEM (using an accelerating voltage of Vacc. ) 75 kV) revealed the presence of ca. 2 nm Rh particles (Figure 1a). For a Ph2PH‚BH3 reaction solution, the presence of small Rh particles was also observed (Figure 1b). However, a solution of the precatalyst only was also found to contain small Rh particles by TEM (Figure 1c). These results suggest that the precatalyst underwent decomposition under the high energy TEM electron beam to form metal particles. Electron beam induced degradation effects such as this are relatively rare but have been previously described in the literature. For example, Schmid has reported that the characterization of gold clusters by high-resolution TEM was hampered by the continual growth of larger clusters as the electron beam damaged the protective ligand shell.16 More recently, high-resolution TEM investigations of copper(I) bisphenanthroline nanoscaffolds have led to the formation of crystalline copper nanoparticles as a result of radiation damage to the supramolecular precursor.17 Fortunately, this electron-beam-induced degradation was preVented when the samples were analyzed by a lower energy electron beam which used an accelerating Voltage of Vacc ) 30 kV. For catalytically active Me2NH‚BH3 solutions, extensive aggregation was observed with the presence of large particles (Figure 1d). Significantly, the analogous case of Ph2PH‚BH3 (Figure 1e) and the precatalyst [{Rh(1,5-cod)(µ-Cl)}2] alone showed only the presence of an amorphous film with no indication of any Rh particles. These results suggest that the catalytic dehydrocoupling of Me2NH‚BH3 is heterogeneous involving Rh metal while that of Ph2PH‚BH3 is homogeneous. (13) Watzky, M. A.; Finke, R. G. J. Am. Chem. Soc. 1997, 119, 10382. (14) In addition to describing nanocluster nucleation and growth, recent work has also described the kinetics of nanocluster aggregation to accurately account for experimental data: Hornstein, B. J.; Finke, R. G. Chem. Mater. 2004, 16, 139. (15) (a) Anton, D. R.; Crabtree, R. H. Organometallics 1983, 2, 855. (b) Whitesides, G. M.; Hackett, M.; Brainard, R. L.; Lavalleye, J. P. P. M.; Sowinski, A. F.; Izumi, A. N.; Moore, S. S.; Brown, D. W.; Staudt, E. M. Organometallics 1985, 4, 1819. (16) (a) Schmid, G. Struct. Bonding 1985, 62, 51. (b) Schmid, G. Chem. ReV. 1992, 92, 1709. (17) Schmittel, M.; Kalsani, V.; Kienle, L. Chem. Commun. 2004, 1534. J. AM. CHEM. SOC.

9

VOL. 126, NO. 31, 2004 9777

ARTICLES

Jaska and Manners

Figure 2. Graph of % conversion vs time for the catalytic dehydrocoupling of Me2NH‚BH3 using [{Rh(1,5-cod)(µ-Cl)}2] (ca. 2 mol % Rh, toluene, 25 °C).

Figure 1. TEM micrographs obtained using Vacc of 75 kV: (a) Me2NH‚ BH3 + [{Rh(1,5-cod)(µ-Cl)}2] reaction mixture; (b) Ph2PH‚BH3 + [{Rh(1,5cod)(µ-Cl)}2] reaction mixture; (c) the precatalyst [{Rh(1,5-cod)(µ-Cl)}2] only. TEM micrographs obtained using Vacc of 30 kV: (d) Me2NH‚BH3 + [{Rh(1,5-cod)(µ-Cl)}2] reaction mixture; (e) Ph2PH‚BH3 + [{Rh(1,5-cod)(µ-Cl)}2] reaction mixture. The white spot in micrograph (e) is a hole in the TEM grid.

Clearly, the use of TEM analysis to provide evidence for the presence of metal particles should be performed with extreme care and with the knowledge that the electron beam may be of sufficient energy to induce degradation of the sample. In particular, when TEM images are used to identify potential active species in catalytic applications, it is highly suggested that control experiments with the precatalyst are performed in order to eliminate the possibility of beam degradation effects that may induce metal nanoparticle formation. Analysis of Catalytically Active Solutions by UV-vis Spectroscopy: For the dehydrocoupling of both Me2NH‚BH3 and Ph2PH‚BH3 with the precatalyst [{Rh(1,5-cod)(µ-Cl)}2], the UV-vis spectra of catalytically active solutions were obtained. For Me2NH‚BH3, a broad plasmon absorption characteristic of Rh colloids was observed (Figure S1, Supporting Information).12 This spectrum was compared against the UV-vis spectrum of the well-defined colloids Rhcolloid/[Oct4N]Cl, which gave a similar broad absorption. However, the spectrum of Ph2PH‚BH3 showed only a large absorption at short wavelengths. The spectrum of the precatalyst [{Rh(1,5-cod)(µ-Cl)}2] also showed a large absorption at short wavelengths with a λmax at 351 nm, which has been previously assigned as metal-to-ligand charge transfer.18 From these spectra, we can conclude that catalytically active Me2NH‚BH3 solutions appear to contain Rh colloids, while those of Ph2PH‚BH3 do not. (18) Epstein, R. A.; Geoffroy, G. L.; Keeney, M. E.; Mason, W. R. Inorg. Chem. 1979, 18, 478. 9778 J. AM. CHEM. SOC.

9

VOL. 126, NO. 31, 2004

Kinetic Curves and Poisoning Experiments for the Catalytic Dehydrocoupling of Me2NH‚BH3 with the Precatalyst [{Rh(1,5-cod)(µ-Cl)}2]: The treatment of Me2NH‚BH3 with a catalytic amount of [{Rh(1,5-cod)(µ-Cl)}2] (ca. 2 mol % Rh) was found to result in a gradual color change from orange to black over about 2 h. With increasing time, the slow formation of a black precipitate and the presence of a “Rh mirror” on the sides of the flask were observed. By monitoring of the dehydrocoupling reaction using 11B NMR, a variable length (45-200 min) induction period was consistently observed.19 Following this induction period, a dramatic surge of catalytic activity was observed resulting in a sigmoidal-shaped kinetic curve (Figure 2). This sigmoidal kinetic curve has been shown to be characteristic for the metal-particle formation reactions A f B (nucleation) and A + B f 2B (autocatalytic surface growth).9,13 Mercury poisoning experiments performed on this catalytic system also supported the presence of heterogeneous catalysis. For dehydrocoupling trials initiated in the presence of excess mercury, no catalytic activity and no color change to black was observed (Figure 3, curve [). When excess mercury was added to a partially dehydrocoupled sample, the catalytic activity was completely suppressed and no further dehydrocoupling was observed (Figure 3, curve 9). In addition, test reactions between mercury and [{Rh(1,5-cod)(µ-Cl)}2] eliminated the possibility of any complicating side reactions that may have led to catalyst deactivation.20 Upon complete conversion of Me2NH‚BH3, the addition of fresh adduct to the catalytically active solution was found to result in immediate dehydrocoupling without an induction period (Figure 4, curve b). When a catalytically active solution was treated with the ligand poison PPh3 (0.5 equiv) and fresh Me2NH‚BH3 added, a substantial reduction in the dehydrocoupling rate was found to result (Figure 4, curve [). The observed partial catalytic activity is likely due to incomplete surface coverage by PPh3, resulting in residual active sites where dehydrocoupling could still occur. Filtration of a catalytically (19) The variability in the length of the induction period may arise from a number of different factors. For example, solvent and Me2NH‚BH3 purity, stirring rate, reaction temperature (inside a glovebox), and the use of different batches of [{Rh(1,5-cod)(µ-Cl)}2] may all affect the length of the induction period observed for a given trial. However, irreproducible induction periods are not unexpected for the in situ formation of a heterogeneous catalyst. See ref 7b for further details. (20) See the Experimental Section for a complete description of the test reactions involving Hg and the precatalyst [{Rh(1,5-cod)(µ-Cl)}2].

Rh-Catalyzed Dehydrocoupling of Borane Adducts

Figure 3. Graph of % conversion vs time for the catalytic dehydrocoupling of Me2NH‚BH3 using [{Rh(1,5-cod)(µ-Cl)}2] (ca. 2 mol % Rh, toluene, 25 °C). At ca. 35% conversion, excess Hg was added to the reaction mixture (curve 9). The dehydrocoupling reaction was initiated in the presence of excess Hg (curve [).

Figure 4. Graph of % conversion vs time for the catalytic dehydrocoupling of Me2NH‚BH3 using [{Rh(1,5-cod)(µ-Cl)}2] (ca. 2 mol % Rh, toluene, 25 °C). Upon complete conversion, Me2NH‚BH3 was added to the catalytically active solution (curve b). Upon complete conversion, the catalytically active solution was treated with 0.5 equiv of PPh3 and more Me2NH‚BH3 was added (curve [). Upon complete conversion, the catalytically active solution was filtered and more Me2NH‚BH3 was added (curve 2).

active solution through a 0.5 µm filter and the addition of fresh Me2NH‚BH3 were found to result in almost complete suppression of catalytic activity (Figure 4, curve 2). All of these tests performed on the Me2NH‚BH3/[{Rh(1,5-cod)(µ-Cl)}2] system suggest the operation of an insoluble, heterogeneous catalyst. Evolution of the Precatalyst [{Rh(1,5-cod)(µ-Cl)}2] during the Dehydrocoupling of Me2NH‚BH3: The decomposition of the precatalyst [{Rh(1,5-cod)(µ-Cl)}2] in the presence of Me2NH‚BH3 was conveniently monitored by 1H NMR. The hydrogenation of precatalyst-derived 1,5-cyclooctadiene to cyclooctane was consistently observed as [{Rh(1,5-cod)(µ-Cl)}2] evolved into the catalytically active Rh colloids during the dehydrocoupling of Me2NH‚BH3. During the induction period (colloid nucleation, from t ) 0-60 min), evidence for cyclooctane or even free 1,5-cyclooctadiene was not detected (Figure S2, Supporting Information). At 15% conversion of Me2NH‚BH3 (t ) 120 min), rapid decomposition of the precatalyst was evident as autocatalytic surface growth occurred. This resulted in nearly 50% conversion of the catalytic amounts of 1,5cyclooctadiene present. At t ) 180 min, 1,5-cyclooctadiene

ARTICLES

conversion was nearly complete (ca. 90%) while Me2NH‚BH3 conversion had reached only 60%. It is important to note that the curve of 1,5-cyclooctadiene conversion also adopted a sigmoidal shape, characteristic of heterogeneous catalysis. These results indicate that the precatalyst underwent rapid decomposition and was fully converted to the active catalyst Rh(0) metal before the dehydrocoupling reaction was complete. Following complete conversion of Me2NH‚BH3, bulk Rh metal was isolated as the active catalyst in ca. 90% yield after removal of all volatile components of the reaction mixture. The resulting black powder was insoluble in common organic solvents such as THF and toluene, but the washings were found to contain [Me2NH2]Cl by 1H and 13C NMR. The presence of [Me2NH2]Cl likely arises from the reaction of free Me2NH (from dissociation of the adduct) and HCl, which may be generated during the reduction step.21 It is likely that small, nanometer sized Rh colloids may be initially formed in solution, with [Me2NH2]Cl acting as an electrostatic stabilizing agent which helps to slow aggregation of the colloids.22 However, evolution of these colloids via aggregation still occurred to form the observed bulk Rh metal, which is a kinetically competent catalyst. For example, use of the isolated active catalyst resulted in dehydrocoupling of Me2NH‚BH3 (ca. 9 h, 25 °C) without a detectable induction period. Catalytic Dehydrocoupling of Me2NH‚BH3 with the (Pre)Catalyst Rh/Al2O3: As the dehydrocoupling of Me2NH‚BH3 can be achieved using a variety of catalysts,4b a selection of other Rh species was subject to the “homogeneous vs heterogeneous” tests. For these catalysts, only the kinetic curves, Hg poisoning, and filtration tests were performed. For Rh/Al2O3 (5 wt. % Rh), the dehydrocoupling of Me2NH‚BH3 proceeded immediately without the presence of an induction period and with high activity (Figure S3, Supporting Information). The lack of an induction period was expected as the catalysis was performed using preformed metal particles attached to a solid support. The addition of excess Hg was found to have no effect on the dehydrocoupling, as poisoning of the active catalyst was not observed. This implies that mercury poisoning of Rh, in this case, does not occur by adsorption to the metal surface but rather by amalgam formation which is more difficult for metals attached to a solid support. Filtration of the solution to remove the insoluble Rh/Al2O3 was found to completely suppress the catalytic activity. These results suggest that Rh/Al2O3 is an insoluble, heterogeneous catalyst for the dehydrocoupling of Me2NH‚BH3. Catalytic Dehydrocoupling of Me2NH‚BH3 with the (Pre)Catalyst Rhcolloid/[Oct4N]Cl: For Rhcolloid/[Oct4N]Cl, the dehydrocoupling of Me2NH‚BH3 proceeded immediately without the presence of an induction period (Figure S4, Supporting Information). However, the addition of excess Hg was found to suppress the dehydrocoupling, as poisoning of the active catalyst was observed. Filtration of the solution through a 0.5 µm filter was found to have only a small effect on the catalytic activity. This was expected as these colloids (ca. 2 nm) are soluble in organic solvents such as THF and toluene, and thus filtration would only remove a minor insoluble component. This (21) The reaction of HCl with Me2NH‚BH3 has been shown to give Me2NH‚ BH2Cl and H2: Jaska, C. A.; Lough, A. J.; Manners, I. Inorg. Chem. 2004, 43, 1090. Therefore, HCl must react with free Me2NH in order to give the observed [Me2NH2]Cl. (22) Roucoux, A.; Schulz, J.; Patin, H. Chem. ReV. 2002, 102, 3757. J. AM. CHEM. SOC.

9

VOL. 126, NO. 31, 2004 9779

ARTICLES

Jaska and Manners

insoluble component may make a small contribution to the overall catalytic activity, as a slight reduction in the rate was observed compared to the unfiltered sample. These results indicate that Rhcolloid/[Oct4N]Cl is a soluble, heterogeneous catalyst23-25 for the dehydrocoupling of Me2NH‚BH3. Catalytic Dehydrocoupling of Me2NH‚BH3 with the Precatalyst [Rh(1,5-cod)2]OTf: For [Rh(1,5-cod)2]OTf, the dehydrocoupling of Me2NH‚BH3 involved the presence of a short induction period (30-45 min) followed by rapid dehydrocoupling to give a sigmoidal-shaped conversion curve (Figure S5, Supp. Info.). A small amount of conversion (